• Nem Talált Eredményt

arXiv:1111.2080v3 [math.PR] 27 Nov 2015

N/A
N/A
Protected

Academic year: 2022

Ossza meg "arXiv:1111.2080v3 [math.PR] 27 Nov 2015"

Copied!
51
0
0

Teljes szövegt

(1)

THE MEASURABLE KESTEN THEOREM

By Mikl´os Ab´ert , Yair Glasner and B´alint Vir´ag Alfr´ed R´enyi Institute of Mathematics , Ben-Gurion University of the

Negev and University of Toronto

We give an explicit bound on the spectral radius in terms of the densities of short cycles in finited-regular graphs. It follows that the a finite d-regular Ramanujan graphGcontains a negligible number of cycles of size less than clog log|G|.

We prove that infinited-regular Ramanujan unimodular random graphs are trees. Through Benjamini-Schramm convergence this leads to the following rigidity result. If most eigenvalues of a d-regular finite graph Gfall in the Alon-Boppana region, then the eigenvalue distribution ofGis close to the spectral measure of thed-regular tree.

In particular,Gcontains few short cycles.

In contrast, we show that d-regular unimodular random graphs with maximal growth are not necessarily trees.

1. Introduction. Let G be a d-regular, finite or infinite connected undirected graph. LetM be the Markov averaging operator on`2(G). When G is infinite, we define the spectral radius of G, denoted ρ(G), to be the norm of M. When G is finite, we want to exclude the trivial eigen- values and thus define ρ(G) to be the second largest element in the set of absolute values of eigenvalues of M. For an infinite graph G, we have ρ(G) ≥ ρ(Td) = 2√

d−1/d where Td denotes the d-regular tree. For fi- nite graphs, the Alon-Boppana theorem [26] says that lim infρ(Gn)≥ρ(Td) for any infinite sequence (Gn) of finite connected d-regular graphs with

|Gn| → ∞.

We callG a Ramanujangraph, if ρ(G) ≤ρ(Td). Lubotzky, Philips and Sarnak [15], Margulis [22] and Morgenstein [25] have constructed sequences of d-regular Ramanujan graphs for d=pα+ 1. Also, Friedman [10] showed that randomd-regular graphs are close to being Ramanujan.

All the Ramanujan graph families above have large girth, that is, the minimal size of a cycle tends to infinity with the size of the graph. How- ever, the reason for that is group theoretic and not spectral, and a priori, Ramanujan graphs could have many short cycles.

MSC 2010 subject classifications:Primary 05C81, 60G50; secondary 82C41

Keywords and phrases:girth, spectral radius, Ramanujan graphs, mass transport prin- cipal, unimodular random graphs

1

arXiv:1111.2080v3 [math.PR] 27 Nov 2015

(2)

In this paper we investigate the connection between the densities of short cycles, the spectral radius and the spectral measure ford-regular graphs. We apply our methods to give explicit estimate these invariants, then we pass to graph limits and prove limiting results.

1.1. Explicit estimates. A cycle (ork-cycle) in a graph is a walk of length k that starts and ends at the same vertex. It is callednontrivial if either for some directed non-loop edge e, the number of times the cycle passes throughediffers from the number of times it passes through the reversal of e, or k = 1 (see Definition 23). For a finite graph G let γk(G) denote the number of nontrivialk-cycles in Gdivided by the number of vertices|G|of G.

Theorem 1. Let G be a finite d-regular graph with |G| ≥ d7. Then for anyk≥1 we have

ρ(G)

ρ(Td) ≥1 +γk(G) νk

3

2log logd−1|G|+ 6 logd−1|G|

where νk= 2·101124k(d−1)3kk.

Applying this to finite Ramanujan graphs yields that they have few cycles of lengtho(log log|G|).

Theorem 2. Let d ≥ 3 and β = (30 log(d−1))−1. Then for any d- regular finite Ramanujan graph G, the proportion of vertices in G whose βlog log|G|-neighborhood is a d-regular tree is at least 1−c(log|G|)−β.

This answers a question of Lubotzky [19, Question 10.7.1] who asked for a clarification on the connection between eigenvalues and girth. Note that until now, it was not even known whether a finite Ramanujan graph cannot have a positive density of short cycles.

It is easy to see that infinite Ramanujan graphs can have arbitrarily many short cycles. In fact, every connected, infinite d-regular graph can be em- bedded as a subgraph of a Ramanujan graph with degree at most d2 (see Corollary 35). However, it turns out that cycles of bounded size must be sparse in a Ramanujan graph.

Theorem3. Let Gbe an infinited-regular graph such that every vertex in Ghas distance at most R from a k-cycle. Then

ρ(G)≥ρ(Td) + d−2

d(d−1)2bR+k/2+1c.

(3)

1.2. Graph limits and spectral measure. The spectral measureµTd of the Markov operator on Td, also known as the Plancherel measure of Td or the Kesten-McKay measure, has density

d 2π

2(Td)−t2 1−t2 .

Let (Gn) be a sequence of finite d-regular graphs. We say that (Gn) has essentially large girth, if for allkthe denisty of nontrivial cycles satisfies

n→∞lim γk(Gn) = 0.

For a finite graph G, let µG denote the eigenvalue distribution of the Markov operator on G. Then the following are equivalent (see Proposition 14):

1. (Gn) has essentially large girth;

2. (Gn) converges toTd in Benjamini-Schramm convergence;

3. µGn weakly converges to µTd.

A sequence (Gn) of finited-regular graphs isweakly Ramanujanif

n→∞lim µGn([−ρ(Td), ρ(Td)]) = 1,

that is, if most eigenvalues ofGnfall in the minimal possible supporting re- gion. Note that a weakly Ramanujan sequence is not necessarily an expander sequence. In fact, the graphsGn do not even have to be connected.

From 1) =⇒3) and the fact thatµTd is continuous, it follows immediately that every graph sequence of essentially large girth is weakly Ramanujan (in contrast,ρis only lower semicontinuous with respect to Benjamini-Schramm convergence of graphs). We show that the converse also holds.

Theorem 4. Let (Gn) be a weakly Ramanujan sequence of finite d- regular graphs. Then (Gn) has essentially large girth.

Theorem4can also be looked at as a rigidity result, as it says that if we force most of the eigenvalues of the Markov operator of a large finite graph inside the Alon-Boppana bound, then their distribution will be close toµTd. In the proof of Theorem4, it is the use of Benjamini-Schramm convergence that allows us to get rid of the bad eigenvalues and clear up the picture. Limit objects with respect to this convergence are random rooted graphs (G, o) called unimodular random graphs. We will sometimes drop the rootofrom the notation. The notion has been introduced in [2]: for the definition, see Section 2. Unimodular random graphs tend to behave like vertex transitive graphs in many senses. Theorem4 now follows from the following.

(4)

Theorem5. Let (G, o)be a d-regular unimodular random graph that is infinite and Ramanujan a.s. ThenG=Td a.s.

This is Kesten’s theorem for vertex transitive graphs ([16] and [28]). We give the following two quantitative versions of Theorem 5. For infinite d- regular unimodular random graphs

(1) Elogρ(G)−logρ(Td)≥ (1

νkk(G, o)

k1Elogκk(G, o).

Here γk(G, o) denotes the number of nontrivial k-cycles starting at o, and νk is a constant defined in Theorem 1. Note that for a fixed finite graph Gthe densityγk(G) equals the expected value ofγk(G, o) over a uniformly chosen rootoof G.

To define κk(G, o), consider all paths of length k from o to a vertex v.

After attaching a fixed path from v to o, these can be used as generators for a random walk on the fundamental group of G. Then κk(G, o) is the geometric average of the spectral radii of these random walks when v is a chosen randomly as the position of the infinite nullcycle (defined in Corollary 20) at timek (see (19), (21) for more details).

Note that if our unimodular random graph G is not a tree, then for k large enough, with positive probability, the Cayley graph of the subgroup of the fundamental group given by above loops as generators has spectral radius less than one. Thus the second bound clearly implies Theorem5.

The first bound in (1) is proved in Section 5 as Theorem 28; the proof uses results from Sections3and 4. It is just the infinite version of Theorem 1. The advantage of this approach is the linear lower estimate on how the spectral radius grows compared to the tree: we believe this to be sharp. The major advantage of the second bound in (1) (proved in Section6) is that it is sharp in limit ask→ ∞, however,κk seems to be hard to compute.

Theorem 4 is related to a paper of Serre [29] that studies asymptotic properties of graph sequences. Let dk(G) denote the number of primitive, cyclically reduced cycles of length k in the graphG. Recall that a cycle is primitive if it is not a proper power of another cycle.

Theorem 6 (Serre). Let (Gn) be a sequence of finite d-regular graphs, such that the limit

γ0k = lim

n→∞dk(Gn)/|Gn|

exists for every k. Then the measures µG weakly converge. If the series

X

k=1

γ0k(d−1)−k/2

(5)

converges then the sequence of graphs is weakly Ramanujan and the limiting measure is absolutely continuous with respect to the Lebesgue measure on [−ρ(Td), ρ(Td)].

Theorem4now immediately implies the following.

Corollary 7. If the series

X

k=1

γ0k(d−1)−k/2

converges, thenγ0k = 0for all kand the limiting measure ofµGn equalsµTd. It is natural to ask whether a version of Theorem 5 holds for growth instead of spectral radius. In Section9we show that the answer is negative:

Theorem8. There exists an infinited-regular unimodular random graph with the same growth as Td but not equal to Td.

We obtain our example by considering the universal cover of the infinite cluster in supercritical percolation overZ2.

1.3. The basic method. There is a common method in the proofs of The- orems1 and 5which we can summarize as follows.

The central tool of our analysis is a nullcycle. Recall that a cycle is a walk of finite length that starts and ends at the same vertex.

Definition 9. A nullcycle is a cycle in a graph G so that if we keep deleting backtrackings (steps that are immediately reversed), we get a cycle of length 0.

This property does not depend on the order of erasing backtrackings.

Equivalently, a nullcycle is a cycle whose lift in the universal covering tree of the graph is also a cycle. In other words, the walk corresponds to a trivial element in the fundamental group of the graphG. The number of nullcycles in ad-regular graph starting at a fixed vertexv equals the number of cycles in thed-regular tree at a fixed vertex.

To bound the spectral radius, we have to count cycles of a given length.

In order to bound the spectral radius away from that of the tree, we need to show that there are exponentially more cycles than nullcycles. Consider the set of cycles of lengthnk starting at a vertexv in a d-regular graphG.

We say that w0 is a rewiring of w if they are at the same place at times

(6)

that are multiples of k. This definition is used in Section 6; in Section 4.2 we use a slight variant of this.

Consider the equivalence class [w] of a typical nullcycle w under the rewiring equivalence relation. The essence of our argument is to show that for a typical w, the probability that a random element C of [w] is a null- cycle is exponentially small. Essentially, in every segment [jk,(j+ 1)k], if there are short cycles around in the graph, there is a positive probability that the rewiring C will use them, and this is likely to stopC from being null-homotopic.

In order to show thatC is nullhomotopic with exponentially small proba- bility, we need to find a linear number ofjso thatGhas short cycles around w(jk). Fortunately, the random nullcyclewsamples the graphGin a homo- geneous manner. In particular, ifw(0) is a uniformly chosen vertex, then so will bew(j) for everyj. This is an advantage of using random nullcycles over random cycles. For infinite graphs, the proof of this step uses unimodularity.

A crucial property that we use to get explicit bounds is one that random nullcycles share with simple random walks. Let G be a d-regular rooted graph and letW be a uniform random nullcycle of length o(p

|G|), starting at the root. Then the expected number of visits of W at any vertex of G can be bounded above in terms ofρ(G) (without referring to the length of the cycle). In particular, for a good expander graph, the expected number of returns of a random nullcycle is bounded. We need this property to show that a typical rewiring will not use the same short cycles over and over again.

This is a technically difficult point that we tackle in Section4.

Putting all these together, we get that if there are many short cycles, then a typical nullcycle will get close to short cycles at linearly many different times. Thus a random rewiring will be a nullcycle only with exponentially small probability. In other words, there are exponentially more cycles than nullcycles, which implies that the spectral radius ofG is greater than that of the treeTd.

1.4. Open problems. It is not clear whether the log log|G|is optimal in Theorem 2. For all the known examples of graphs that are close to being Ramanujan, the shortest cycles with positive density are actually logarith- mic.

Problem 10. Is there a constant c = c(d) > 0 such that for any d- regular Ramanujan graph sequence (Gn), the probability that the clog|Gn|- neighborhood of a uniform random vertex in Gn is a tree converges to 1?

A standard ergodicity argument says that for an ergodic unimodular ran-

(7)

dom graph G, the weak limit of the random walk neighborhood sampling of G gives back the distribution of G a.s. [6]. This suggests the following possible generalization of Theorem 5.

Problem 11. Let G be an infinite d-regular rooted Ramanujan graph and let k >0. Let pn denote the probability that the random walk of length non G ends on a k-cycle. Is it true thatpn converges to 0?

That is, is it true that the random walk neighborhood sampling ofGcon- verges to Td? The answer does not follow from Theorem 5, even when the random walk sampling converges, as the limit is only a stationary distribu- tion on rooted graphs and is not necessarily unimodular. It would also be interesting to see whether Theorem5 holds for stationary random graphs.

The recent paper [14] solves Problem 11 affirmatively in the case when the so-called co-growth of G, the exponent of the probability of return for a non-backtracking random walk, is less than 1/√

d−1. However, when the co-growth equals 1/√

d−1, the graph is still Ramanujan but the answer seems unclear. We thank Tatiana Smirnova-Nagnibeda for communicating this with us. After the first preprint version of this paper appeared, R. Lyons and Y. Peres, [20] generalized our results and in particular gave a positive answer to Problem11.

The linear lower estimate in the spectral radius in Theorem 1 seems to be sharp, but we have not been able to settle this with a suitable family of examples. The same is true for unimodular random graphs (see the first bound of (1)).

Problem 12. Does there exists C > 0 such that for every r > 0 there exists an infinited-regular unimodular random graph Gwith

ρ(G)≤ρ(Td) +Cr such that the density of loops in G is at leastr?

One natural idea would be to use a modified universal cover of a finited- regular graph of sizenwith a loop, where we never open the loop in the cover.

It looks reasonable that this cover (which is a finitely supported random rooted tree with loops) should have spectral radius aroundρ(Td) +C/n.

The paper is organized as follows. Section2contains the basic definitions and we prove some lemmas that will be used throughout the paper. In Sections 3 and 4 we use properties of cycles in trees to study nullcycles,

(8)

which are needed for Theorem 1. In Section 5 we prove Theorem 28, a more general version of Theorems 1 and 5. We also show Theorem 29, a more general version of Theorem 2. Finally, in this Section we also prove Theorem4.

Section6contains a sharp bound on the spectral radius in terms of random walks on the fundamental group. Section 7 contains the proof of Theorem 3. This section is independent of the rest.

In Section8we give example of Ramanujan graphs with loops, and Section 9we prove Theorem 8.

Note that one can read Section 5, and 6 independently, after reading Section 2, but reading any of these two will give help when reading the other.

An earlier version of this paper contained a generalization of Kesten’s theorem on groups. As the readership of this result is expected to be different from that of the current paper (and the current paper is already long), we decided to publish it in a separate article, see [1].

2. Preliminaries . In this section we define the notions and state some basic results used in the paper.

We follow Serre’s notation for a graph, with a modification on how to define loops. A graph G consists of two sets, a set of vertices denoted by V(G) and a set of edges denoted E(G). For every edge e ∈ E(G) there are vertices e (the initial vertex) and e+ (the terminal vertex). We allow e = e+: such edge is called a loop. For every edge e there is a reverse edge e ∈ E(G) such that e+ = e and e = e+. For a loop e, we allow e=e; these are calledhalf-loops. The degree of a vertexv is

degv=

e∈E(G)|e=v

So half-loops contribute 1 to the degree, but loops together with their dis- tinct inverse contribute 2. For spectral and random walk questions, each (non-half) loop can be replaced by two half-loops. So in this paper we will assume that all loops are half-loops.

A graph isd-regular, if all vertices have degreed.

A walk of length n is a sequence of directed edges w= (w1, w2, . . . , wn) such that w+i−1 = wi (2 ≤ i ≤ n). The walk is a cycle if w1 = wn+. The vertices of the walk are defined by w(i−1) = wi and w(n) = w+n is the end of the walk. The inverse of a walk w is defined by w−1 =

(9)

(wn, wn−1, . . . , w1). A cycle is a nullcycleif its lift to the universal cover of Gstays a cycle. That is the same as saying that if we keep erasing backtracks from the cycle, we get to the empty walk. For a rooted graph (G, o) we will denote the set of nullcycles of lengthnby Nn.

For a graph G and x, y ∈ V(G) let Wn(x, y) denote the set of walks of length n starting at x and ending at y. A random walk of length n starting at x is a uniform random walk starting at x. Let pn(x, y) denote the probability that a random walk of length nstarted at x ends at y. We callpn(x, x) then-stepreturn probability.

LetGbe ad-regular, connected undirected graph. Let`2(G) be the Hilbert space of all square summable functions on the vertex set ofG. Let us define the Markov operatorM :`2→`2 as follows:

(M f)(x) = 1 d

X

e∈E(G),e=x

f(e+)

When G is infinite, we define the spectral radius of G, denoted ρ(G), to be the norm of M. When G is finite, we want to exclude the trivial eigenvalues and thus defineρ(G) to be the second largest element in the set of absolute values of eigenvalues ofM. Note that when the connected graph G is bipartitie, then −d is an eigenvalue with multiplicity one; this is not counted in the definition ofρ(G).

In the case whenGis infinite and connected, one can express the spectral radius of Gfrom the return probabilities as follows:

ρ(G) = lim

n→∞p2n(x, x)1/2n wherex is an arbitrary vertex ofG.

The Markov operator M is self-adjoint, so we can consider its spectral measure. This is a projection valued measure P such that P(O) : l2(G)→ l2(G) is a projection for every Borel set O ⊂ [−1,1]. For every f ∈ l2(G) withkfk2 = 1, the expression

µf(O) =hP(O)f, fi defines a Borel probability measure on [−1,1].

For graphG rooted ato, let the spectral measure of G be µG,oδo

where δo ∈ l2(G) is the indicator function of o. The best way to visualize this measure is to look at its moments, that satisfy the following equality:

Z

[−1,1]

xkG,o=pk(o, o)

(10)

for all integersk≥0.

Unimodular random graphs. Heuristically, a unimodular random graph is a probability distribution on rooted graphs that stays invariant under mov- ing the root to any direction. However, one has to be careful with this intu- ition, as direction is not well-defined and indeed, there exist vertex transitive graphs that we want to exclude from the definition. We follow [3, Section 5.2] in our definition restricted to the d-regular case where it is somewhat simpler.

A flagged graph is a graph with a distinguished root and a directed edge starting at the root. One can invert the flag by moving the root to the other end of the flag and switching the direction of the flag.

Definition 13. Let G be a probability distribution on rooted d-regular graphs. Pick a uniform random edge from the root and put a flag on it. This gives a probability distribution Ge on flagged d-regular graphs. We say that G is a unimodular random graph, if the distribution Ge stays invariant under inverting the flag.

That is, if some of the flagged lifts of a given rooted graph are isomorphic, we count it with multiplicity. Note that vertex transitivity in itself does not imply unimodularity. A simple example is the so-called grandmother graph. This can be obtained by taking a 3-regular tree and directing it towards a boundary point, then connecting every vertex to the ascendant of its ascendant (its grandmother) and then erasing directions (see Figure1).

If one does not mind working with edge directed graphs, it is easier to see the lack of unimodularity in the oriented 3-regular tree itself. There is only one type of rooted graph here that obviously appears with probability 1.

The corresponding measure on flagged graphs puts the flag on an outgoing edge with probability 1/3, but after an inversion we see an outgoing edge with probability 2/3. See [2] for more about unimodularity.

Fig 1. The grandmother graph

(11)

Mass Transport Principle. The most useful property about unimodular random graphs (that can also be used to define them) is the Mass Transport Principle which is as follows. Let f be a non-negative real-valued function on triples (G, x, y) where G is a d-regular rooted graph and x, y ∈ G such thatf does not depend on the location of the root. Then the expectations

E

 X

y∈G

f(G, o, y)

=E

"

X

x∈G

f(G, x, o)

#

whereois the root of G. The picture is that if one sets up a paying scheme on the random graph G that is invariant under moving the root, then the expected payout of the root equals its expected income.

Benjamini-Schramm convergence. Ad-regulargraph sequence(Gn) is defined as a sequence of finited-regular graphs with size tending to infinity.

By a pattern of radius r we mean a rooted graph where every vertex has distance at most r from the root. For a finite graph G and a pattern α of radius r let the sampling probability p(G, α) be the probability that ther-ball around a uniform random vertex of Gis isomorphic to α. We say that a graph sequence (Gn) is Benjamini-Schramm convergent, if p(Gn, α) is convergent for every patternα. It is easy to see that every graph sequence has a convergent subsequence.

What is a natural limit object of a convergent graph sequence? One can also take pattern densities of a unimodular random graphG; therep(G, α) denotes the probability that the r-ball around the root of G is isomorphic toα. We say that a graph sequence (Gn) converges toGif

n→∞lim p(Gn, α) =p(G, α) for all patternsα.

Every Benjamini-Schramm convergent graph sequence has a unique limit unimodular random graph (see [3, Section 2.4]).

For a finite d-regular graph G let µG denote the eigenvalue distribution of the Markov operator onG. Note that for a uniform random vertexo we have µG=EµG,o. For an infinite unimodular random graph G we can also defineµG=EµG,o.

Proposition14. Let(Gn)be a sequence of finited-regular graphs. Then the following are equivalent:

1) (Gn) has essentially large girth;

2) (Gn) converges toTd in Benjamini-Schramm convergence;

3) µGn weakly converges to µTd.

(12)

Proof. The equivalence of 1) and 2) is immediate from the definition of Benjamini-Schramm convergence.

Assume that (Gn) converges to the unimodular random graph G. We claim that µGn weakly converges to the expected spectral measure µG = EµG,o. To check this, we can look at the kth moment

Z

xkG =E

pGk(o, o) .

Recall that pGk(o, o) denotes the probability of return of the random walk on G starting at o. But for any graph G and vertex v of G, the return probabilitypGk(v, v) only depends on thek/2-ball aroundo. Since there are only finitely many patterns of a given radius, this implies

E

pGk(o, o)

= X

αis a pattern of radiusbk/2c

p(G, α)pαk(v, v) wherev is the root ofα. Now (Gn) converges toG, so

E

pGk(o, o)

= lim

n→∞

X

αis a pattern of radiusbk/2c

p(Gn, α)pαk(v, v) =

= lim

n→∞E h

pGkn(u, u)i

= lim

n→∞

Z

xkGn

whereuis a uniform random vertex inGn. So,µGn weakly converges toµG as claimed. Hence 2) implies 3) follows immediately.

Assume that 1) does not hold, that is, (Gn) is a graph sequence that does not have essentially large girth. Then there exists k, ε > 0 such that the density of k-cycles in Gn is at least ε for infinitely many of the Gn. This implies that for thesen,

Z

xkGn =Eh

pGkn(u, u)i

≥pTkd(o, o) + ε dk =

Z

xkTd+ ε dk

which implies thatµGn does not converge weakly toµTd. Hence, 3) does not hold. We proved the required equivalences.

Fundamental group. Let G be a graph rooted at o. We call two cycle starting atohomotopic, if one can get one from the other by inserting and erasing backtracks, that is, walks of typess where sis an edge of G. Then the set of equivalence classes forms a group under concatenation, called the fundamental groupπ1(G). It is well known that the fundamental group of a graph without half-loops is a free group [24, Theorem 5.1]. Every half-loop

(13)

adds a cyclic group of order 2 as a free product. The most important general property of fundamental groups we shall use in this paper is that if H is a subgraph of G, then the induced homomorphism from π1(H) toπ1(G) is injective.

3. Cycles inTd. This section establishes some basic properties ofNn= Nn(d), the set of n-cycles in Td. Such a cycle α ∈ Nn in the 3-regular tree is depicted in Figure 2. Given any covering map p:T3 → X to a 3-regular graphX, the projection of the cyclep(α) is referred to as a null cycle in the graphX.

T3

p

X

α

p(α)

Fig 2. A cycle in the3-regular tree

3.1. Explicit return probability bounds. We start by estimating the size ofNn.

Lemma 15 (Return probabilities of SRW on Td). Let ρ = ρ(Td) = 2√

d−1/d. The n-step return probability rn = d−n|Nn(d)| for simple ran- dom walk in Td for evenn >0 satisfies

2 3

ρn

n3/2 < rn<10 ρn n3/2.

(14)

Proof. Return probabilities are moments of the spectral measure. The spectral measure inTd is supported on [−ρ, ρ] with density given by

d 2π

2−t2 1−t2 ,

see [31], formula (19.27). So for even n, by symmetry, we may write rn= d

π Z ρ

0

tn

2−t2

1−t2 dt= d 2π

Z ρ2 0

s(n−1)/2

2−s 1−s ds.

Then, with

a=ρ−2 Z ρ2

0

s(n−1)/2p

ρ2−s ds=

√π

2 ρnΓ(n/2 + 1/2) Γ(n/2 + 2) we have

2

2π a < rn< dρ2 2π(1−ρ2) a.

A small computation shows that ford≥3 we have 8

3 ≤4− 4

d =dρ2, dρ2

1−ρ2 = 4d2−4d (d−2)2 ≤24.

Now forn≥4 we have

κn−3/2 ≤ Γ(n/2 + 1/2)

Γ(n/2 + 2) ≤23/2n−3/2, κ= 43/2Γ(2.5) Γ(4) .

The upper bound also holds for n= 2. (We manually check that the lower bound of the lemma holds forr2 = 1/d.) To complete the proof, we bound the lower and upper constants factors

8 3

1 2π

√π 2 κ= 2

3, 9.57∼24 1 2π

√π

2 23/2= 12p

2/π <10.

Our next goal is to study the expected number of visits for random cycles inTd. This will be based on the same question for random walk excursions onZ. Recall that anexcursionof lengthnonZis a walk that stays positive except for time 0 andn, when it is zero.

(15)

3.2. Visits of cycles.

Lemma 16 (Counting excursions). Let wn,k be the number of walks of length n≥1 from 0 to k≥0 in Z. Then

wn,k<p 2/π 2n

√n e−k2/(2n).

Let wn,k+ be the number of such paths that stay positive after time 0. Then for k >0 we have

w+n,k<p

2/π2nk

n3/2 e−k2/(2n).

Proof. We may assume that nand kare the same parity. Then wn,k=

n

n+k 2

.

We use the inequality

n bn/2c

<p 2/π 2n

√n,

which holds since the ratio of the two sides is increasing along even (respec- tively odd)n and converges to 1. Forn even we now write

n

n+k 2

n n/2

−1

= ((n−k)/2 + 1)· · ·(n/2) (n/2 + 1)· · ·((n+k)/2)≤

n−k n

k/2

≤e−k2/(2n),

and the odd case follows similarly.

By the Ballot theorem (see Section 2.7.1 in [18]) we have w+n,k= k

nwn,k ≤p

2/π2n k

n3/2 e−k2/(2n).

Recall that a simple random walk excursion of length n on Z is a uniform choice from all excursions of length n. In other words, it is the simple random walk conditioned to stay positive except for time 0 and n, when it is zero. Now we are ready to bound the expected number of visits for simple random walk excursions on Z.

Lemma 17 (Visits of SRW excursions on Z). The expected number of visits vk,n to level k > 0 for the simple random walk excursion of length n onZ satisfiesvk,n ≤64k.

(16)

Proof. Let wn,k+ denote the number of walks of length n starting at 0 and ending at k ≥0 that stay positive except perhaps at time 0 and n. If Xm is a random walk excursion of lengthn, then

vk,n=E

n−1

X

m=1

1(Xm =k) =

n−1

X

m=1

P(Xm =k) = 1 wn,0+

n−1

X

m=1

wm,k+ wn−m,k+ ≤ 2 w+n,0

n/2

X

m=1

w+m,kw+n−m,k

For n = 2 the claim is easy to check. For n ≥ 4 even we have the lower bound using the Catalan number formula

w+n,0 = 2wn−2,0

n ≥ 1

√2π 2n n3/2,

where the last inequality holds since the ratio of the two sides is decreasing and converges to 1. Together with Lemma16 this gives the bound

vk,n≤2·2 πk2

2πn3/2

n/2

X

m=1

e−k2/(2m)

m3/2(n−m)3/2 ≤ 2 π·2·√

2π·23/2k2

n/2

X

m=1

e−k2/(2m) m3/2 . Letam denote the last summand, even for non-integerm. Then for allm≥1 and δ∈[0,1] we haveam+δ ≥2−3/2am. Thus we can bound the sum by

23/2 Z

1

e−k2/(2x)

x3/2 dx <23/2 Z

0

e−k2/(2x)

x3/2 dx= 4√ π k .

Arandom cycleis a cycle chosen from uniform measure from the set of cycles with the same starting point.

Lemma18 (Visits of cycles in Td). The expected amount of time a ran- dom cycle of even lengthn in Td spends at distance k >0 from its starting point is at most 2·104k. Fork= 0 it is at most 301.

Proof. Consider a random cycle of length ninTd from the root o. Let Rj be the distance of the walk fromoat timej. The following is well-known, see Section 2 of [8].

Let 0 =T0 < T1<· · ·< TM =nbe the (random) times when Rj is zero.

Given the values ofTi andM, the sections ofRj in between are independent simple random walk excursions onZ. In particular, given this information, Lemma17 implies that the conditional expectation of the number of visits ofRj tok is bounded above by 64kM. So by Lemma17 it suffices to show thatEM is bounded by a constant independent ofn.

(17)

Let rn be the probability that the simple random walk on Td visits its starting point at timen. By the Markov property, we have

EM = 1+

n/2−1

X

k=1

P(R2k= 0) = 1+1 rn

n/2−1

X

k=1

r2krn−2k≤1+3 2·102

n/2−1

X

k=1

n3/2 (2k)3/2(n−2k)3/2 where the last inequality follows form Lemma 15. Since the summand is

convex as a function ofk, thekterm is bounded above by Z k+1/2

k−1/2

n3/2

(2x)3/2(n−2x)3/2 dx and the entire sum is at most

Z n/2−1/2 1/2

n3/2

(2x)3/2(n−2x)3/2dx= 2(n−2) p(n−1)n <2 This givesEM <301.

Finally, we consider the limiting process of the random cycle starting at oinTd.

Proposition 19 (The infinite cycle in Td). Let (Xkn, k = 0. . . n) be the random cycle of even length n from o to o in Td. Then as n→ ∞ the process (Xkn, k = 0. . . n) converges in distribution to a process (Xk, k ≥0) called the infinite cycle, a time-homogeneous Markov process with transition probabilities (2).

Proof. The random cycle of length n is a time-inhomogeneous Markov process. Letpnk(x, y) be denote its transition probabilities fromxtoyat time k. It suffices to show that the ratios ofpnk(x, x+)/pnk(x, x) converge, (where x+, x denotes a child or the parent of x, respectively) as any probability of the form

P(X1n=x1, . . . , Xkn=xk)

can be written as an expression containing finitely many of these probabili- ties. Withpn(x, y) denoting the simple random walk transition probabilities inTd, the standard path counting argument gives

pnk(x, x+)

pnk(x, x) = pn−k−1(x+, o) pn−k−1(x, o).

(18)

we now use Theorem 19.30 in [31] which for x fixed andn→ ∞ gives pn(o, x) = (c+o(1))

1 +d−2 d |x|

(d−1)−|x|/2ρ(Td)nn−3/2 where|x|is the graph distance ofx from o, to get

(2) lim

n→∞

pn−k−1(x+, o) pn−k−1(x, o) = 1

d−1

d+ (d−2)(|x|+ 1)

d+ (d−2)(|x| −1) =: p(x, x+) p(x, x). So (Xk, g≥0) is a time-homogeneous Markov process with transition prob- abilitiesp (which are determined by (2) since they sum over the neighbors of x to 1). Clearly|Xn|is also a time-homogeneous Markov process, which has up/down transition probability ratio from x∈Z+ given by

d+ (d−2)(x+ 1) d+ (d−2)(x−1).

Note that whend= 2 we get the reflected simple random walk, as expected.

Corollary20 (The infinite nullcycle). Let Gbe ad-regular graph, and ( ¯Xkn, k= 0. . . n)be thekth step of a uniformly chosen random nullcycle from a vertex o to o. Then X¯kn converges in distribution as n→ ∞ to a limiting process( ¯Xk, k≥0)called the infinite nullcycle. In particular, the fixed-time distributions converge.

Proof. Note that ¯Xkn is just the image under the universal cover map fromTdtoGof the random cycle inTd. So the claim follows from Proposition 19.

4. Properties of nullcycles. This section establishes some important properties of random nullcycles in graphs. But first we need a simple well- known lemma.

Lemma 21 (Spectral radius and hitting probabilities). Let G be a con- nected d-regular graph and let o be a vertex. Let pn(o, A) denote the proba- bility that a random walk of length n starting at o ends in the finite vertex setA. Then with the spectral radius ρ(G) we have

pn(o, A)≤p

|A|ρ(G)n+2|A|

|G|.

(19)

Proof. We prove the claim for finite graphs, the infinite case is similar but simpler. Let m =|G|, the number of vertices of G. Let v0 denote the function on G that takes value 1/√

m everywhere. Then v0M = v0. When Gis not bipartite, let l2(G) denote the orthogonal subspace of v0 inl2(G).

When Gis bipartite, let I be an independent subset of Gof size m/2 con- tainingoand letv1 be the function onGthat takes values 1/√

mon I and

−1/√

m otherwise. Then v1Mn = (−1)nv1. Let l2(G) denote the subspace orthogonal to v0 and v1 inl2(G).

Now ρ(G) equals the norm of M on l2(G). Let δA denote the indicator function of the vertex setA. Letv be a projection ofδo onto l2(G), and let vo−v. Then kvk ≤ 1. ForG bipartite, we can write v =a(v0+v1), witha= 1/√

m. We have

(3) hvMn, δAi=ha(v0+v1)Mn, δAi=ha(v0+ (−1)n)v1, δAi

Sincev0and v1 are orthonormal, writingδAin the orthonormal basis we see that (3) is bounded above by

ahv0, δAi+a|hv1, δAi| ≤2|A|/m.

Similarly, in the non-bipartite case hvMn, δAi=|A|/m. We now have

pn(o, A) =hδoMn, δAi=hvMn, δAi+hvMn, δAi ≤2|A|/|G|+kvk ·ρ(G)n· kδAk.

HerekδAk=p

|A|. The claim follows.

4.1. Visits of nullcycles.

Proposition22 (Visits of nullcycles). For any infinited-regular rooted connected graph (G, o) with ρ(G) < 1 the number of visits VA to a finite vertex set A of a random nullcycle of length nstarting at osatisfies

EVA≤2·104 |A|

(1−ρ(G))2.

This is at most 107|A|if ρ(G)≤19/20. Note that19/20> ρ(Td) for d≥3.

For any finite d-regular graph G we also have

EVA≤4·104|A|

1

(1−ρ(G))2 +72n2

|G|

.

This is at most2·107|A|if ρ(G)≤19/20 and n2 ≤ |G|.

(20)

Proof. LetXj be a random cycle in thed-regular treeTdstarted at the rooto, and let ¯Xj be its projection to the graphG. Then we have

EVA=E

n

X

j=0

1( ¯Xj ∈A) =

n

X

j=0

P( ¯Xj ∈A).

Condition on|Xj|, the distance from the root, and then sum over all possible options to get

EVA=

n

X

j=0 n

X

k=0

P(|Xj|=k)P( ¯Xj ∈A| |Xj|=k).

Note that given|Xj|=k, the distribution ofXj is uniform on the k-sphere aboutoin the tree. Thus the distribution on ¯Xj in the graphGis that of the kth step of a nonbacktracking random walk. So letpkdenote the probability that thekth step of the nonbacktracking walk is in A.

Switching the order of summation we get EVA=

n

X

k=0

pk

n

X

j=0

P(|Xj|=k)≤500p0+

n

X

k=1

2·104kpk

where the last inequality is based on the fact that the j-sum gives the ex- pected number of visits to distance k for the random cycle in Td, and the result of Lemma 18. Note thatp0 = 1(o ∈A). The above can be bounded by Green function techniques as follows. Define

C(z) =

X

k=0

pkzk,

the generating function for the proportion of nonbacktracking paths that start from oand end in A. For anyz∈(0,1] we have

n

X

k=0

kpk≤z1−n

X

k=1

kpkzk−1 =z1−nC0(z)

The right hand side is a power series with nonnegative coefficients, so it always makes sense but may equal +∞. Rewriting our bound in terms of C we get

EVA≤2·104z1−nC0(z) + 500·1(o∈A).

(21)

Let G(z) be the analogous generating function for simple random walk. It was shown in [5], (see formula (2.3) in [27]) that for anyd-regular graph we have

C(z) = 1(o∈A)

d + (d−1)2−z2 d(d−1 +z2) G

d z d−1 +z2

. Now withx=dz/(d−1 +z2) we compute

C0(z) =a0G(x) +a1G0(x).

where

a0 = − 2(d−1)z (d+z2−1)2 ≤0, a1 = d3−d2 z2+ 3

+d z2+ 3

+z4−1 (d+z2−1)3 ≤1,

for our range of parametersd≥2 andz∈(0,1]. We now consider two cases.

1. ForG infinite with ρ =ρ(G) <1, we use the case z = 1, noting that the radius of convergence of G is 1/ρ > 1. Since G and its derivative are nonnegative, we get the upper bound

1

210−4EVA≤ |A|+G0(1)≤ |A|+ ¯G0(1), G(z) =¯ |A|

1−zρ.

The last inequality uses the fact that the probability that simple random walk at time k is in A is bounded above by |A|ρk, so we can replace G0(z) by ¯G0(z). Finally, we have

|A|+ ¯G0(1) =|A|1−ρ+ρ2

(1−ρ)2 ≤ |A|

(1−ρ)2.

2. For G finite, we use the case z < 1. Since G and its derivatives are nonnegative, we get the upper bound

C0(z)≤ G0(x)≤G¯0(x).

For the last inequality, we useρ=ρ(G), G(x) =¯ |A|

X

k=0

xkk+ 2/|G|) = 2

|G|

|A|

1−x + |A|

1−xρ. and use Lemma21 to bound the return probabilities. This gives

|A|+ ¯G0(x) = 2|A|

|G|

1

(1−x)2 +|A|ρ+ (1−ρx)2 (1−ρx)2 ≤ 2

|G|

|A|

(1−x)2 + |A|

(1−ρ)2.

(22)

We now have 1

1−x = d−1 +z2

(d−1−z)(1−z) ≤ d d−2

1

1−z ≤ 3 1−z and setz= 1−1/(2n) to get

EVA≤2·104z−n(C0(z)+|A|)≤2·104(1−1/(2n))−n|A|

2·32·22n2

|G| + 1

(1−ρ)2

since forn≥1 the (1−1/(2n))−n≤2, and the claim follows.

4.2. Cycles and nullcycles. We now turn to the connection between or- dinary cycles and nullcycles. We recall the definition of nontrivial cycles.

Definition 23. Call a cycle of length kin a graph a nontrivial cycle if either

• for some directed non-loop edgee, the number of times the cycle passes throughediffers from the number of times it passes through the reversal of e

• or k= 1.

This definition differs slightly from “vanishing in homology”, but is pre- cisely what we need in our proof (briefly we useZ-homology fork≥2, and Z2-homology for k = 1). Our goal there is to take a nullcycle and make it non-nullhomotopic. We do this by swapping the direction of nontrivial sub-cycles of length k≥2. For loops this does not work (they do not have direction), so we have to have a separate argument for k = 1: we add or erase them.

Cycles not covered by this definition are calledtrivial. For example, null- cycles are trivial and simple cycles are nontrivial.

The following theorem is another main ingredient in the proof of Theorem 1. LetNndenote the set of nullcycles starting atoin the rooted graph (G, o).

Theorem 24 (Cycles and nullcycles). Let (G, o) be a d-regular rooted graph, and let n, k, ` >0.

For a nullcyclew ∈ Nnk let χ(w, a, k) =χ`(w, a, k) denote the indicator function that the path segmentwa, . . . , wa+k is a nontrivial k-cycle and that the vertexwa is visited at most ` times by w. Let

(4) χw =

n−1

X

j=0

χ(w, jk, k).

(23)

Then with c1 = 1/16 and ck = (d−1)−k/2 (for k≥2) we have

|Wnk(o, o)| ≥ 1 14

X

w∈Nnk

exp (ckχw/`), where Wnk(o, o) is the set of cycles of length nk starting at o.

The quantity χω will be estimated in terms of the parameters γk(G).

Heuristically, if it is large, it means that there are many different places in ω where rewiring is possible. The point in limiting the number of visits by

`is a convenient way to make sure that if there are many possible rewiring times, then they happen also at many different locations.

Proof. Let us denoteW =Wnk(o, o), andN =Nnk, the subset of null- cycles. We first break W into equivalence classes, called rewiring classes.

A loop is calledsingle if its vertex has no other loops. Otherwise, we call it amultiple loop.

When k = 1 we break up the sum on the right of (4) into a sum over single loops and a sum over multiple loops, counted asχ1w2ww. We choose k (and for k = 1 we choose single or multiple loops), and consider rewiring classes depending on our choice.

Casek= 1, single loops. Given a pathw, let ¯wdenote the path in which all self-loops whose vertex is visited at most`times (not counting consec- utive visits) have been erased. Letw ≡w0 if ¯w= ¯w0. (“Not counting consecutive visits” means that visits tovthat are at consecutive times count as a single visit.)

Casek= 1, multiple loops. Two paths are equivalent if for all times ithe vertices satisfy wi = w0i, and w and w0 agree except at times when they traverse multiple self-loops.

Casek≥2. The pathsw andw0 are equivalent if for all 0≤j ≤n−1 the following holds

• Ifwjk 6=wjk+k then the path segment between these times of w andw0 is equal.

• If wjk =wjk+k and the path segment between these times of w is trivial, then it equals the corresponding path segment inw0.

• Ifwjk =wjk+kand the path segment between these times ofwis nontrivial then it either equals the corresponding path segment inw or is the time-reversal of that. We calljk a proper cycle timeofw, and the corresponding path segment aproper cycle ofw.

(24)

This is illustrated in the example depicted in figure3.

Forw∈W let [w] denote the equivalence class ofw, called rewiring class.

Note that the rewiring defined here is more complex than the one in Section 1.3. For w ∈ N let p(w) denote the probability that a uniform random element of [w] is nullhomotopic.

Then we have

|W|= X

Ais a rewiring class

|A| ≥ X

Ais a rewiring class,A∩N 6=

|A|= X

w∈N

|[w]|

|[w]∩ N | = X

w∈N

p(w)−1.

What remains is to show that for allw∈ N we have p(w)≤14 exp(−ckχw/`).

We will do this case by case.

Casek= 1, single loops.We call a vertex with a single loop (and its loop) reclusive for w, if w visits it at most ` times (not counting consecutive visits). Whether a vertex is reclusive or not depends only on [w].

Let τi, i = 1, . . . , κ denote the times when ¯w visits a reclusive vertex, and letφbe the number of loops erased from w to get ¯w. Then an element of [w] is determined by X1, . . . Xκ, the number of loops inserted into ¯w at timesτ1, . . . , τκ. A uniform random element of [w] corresponds to a uniform random choice of theXi so that their sum isφ. Let trwdenote the function that assigns to every reclusive loop of [w] the number of times modulo 2 thatw passes through it. Then

p(w)≤P(trw= 0),

where the right hand side refers to a random element of [w]. This is exactly the probability that for each reclusive vertex the sum of theXicorresponding to that vertex is even. By Lemma25this is at most 14 exp (−min(m, φ/`)/14), where m is the number of different reclusive vertices visited. Note that m≥χ1w/`and φ≥χ1w, so we get the bound 14 exp −χ14`1w

.

Case k= 1, multiple loops. We call a vertex importantif it has a loop traversed byw. Further, we call a loop important if its vertex is important (even if not traversed byw). Note that the set of important loops (or vertices) only depends on the equivalence class ofw.

For a path, let tr denote the function that assigns to each important loop the number of times modulo 2 that it is traversed. Consider a random element w of [w]. For each important vertex v with kv loops, let ¯Xv =

(25)

(Xv,1, . . . , Xv,kv) record the number of times w visits its loops. Note ¯Xv

are independent asv varies, and each have a multinomial distribution with probabilities 1/kv for each option; each traversal is assigned to one of the loops uniformly at random.

Conditioning on the assignment of all traversals except for the last one, we see that the probability that Xv,1, . . . , Xv,kv are all even numbers is at most 1/kv ≤1/2. So ifiis the number of important vertices visited at most

`times, then we havei≥χ2w/`and

p(w)≤P(trw= 0)≤2−i ≤2−χ2w/`.

Casek≥2.For a path, let tr denote the antisymmetric edge function that sums 1 over all forward steps of a path and−1 over all backward steps (here ignoring self-loops). Note that the trace of a random element w in [w] can be written as

(5) trw=trw¯+ X

proper cyclescofw

Xctrc

where theXc are independent random variables uniform on{−1,1}, and ¯w denoteswwith all its proper cycles removed. We claim that

p(w)≤P(trw= 0)≤2−|w|o

where|w|o is the maximum size of a subset of linearly independent proper cycles of w. Indeed, consider such a set C, and complete it to a basis for antisymmetric edge functions. Fix all values ofXcforc /∈ C. Then forc∈ C, looking at the ac-coordinate of the equation (5), we see that it can hold only ifXcequals some fixed value, which has probability 1/2 or 0, independently over the coordinates. The claim follows.

Our next step is to bound the number of independent cycles. Fix aj0, and we consider the setJ of indicesj so that theχ(w, jk, k) =χ(w, j0k, k) = 1, and the cycles of w at jk and j0k share an edge. For a vertex v let J(v) denote the number ofj∈J so that wjk =v. Since for j∈J the vertexwjk

is visited at most `times, we have J(v) ≤`. If two k-cycles share an edge, then a vertex on one and a vertex on the other are of distance at mostk−1 from each other. Thus we have

|J|= X

v∈B(wj0k,k−1)

J(v)≤`|B(wj0k, k−1)| ≤`d(d−1)k−2,

whereB(v, r) is the ball of radiusraboutv. This means that the dependency graph of cycles has degree at most d(d−1)k−2`and size χw, and therefore

(26)

w0k,10k

w9k

w1k,2k,3k

w5k

w8k

w4k

w6k

w7k

This path has one independent proper cycle.

k= 3

Fig 3. A null cycleω∈ N30 with two proper cycles of lengthk= 3. These are opposite to each other and in particular dependent. Changing the direction in one of them gives raise to a nontrivial cycle equivalent toω.

contains an independent set of sizeχw/(d(d−1)k−2`+ 1). So we getp(w)≤ 2−χw/(d(d−1)k−2`+1) ≤e−χw/(2(d−1)k`).

Now we have eitherχ1w78χw orχ2w18χw. In either case, we get p(w)≤14 exp(−χw/(16`)).

Together with thek≥2 case, this completes the proof.

The following simple probabilistic lemma was used in the proof of Theo- rem24.

Lemma25. Let X = (X1, . . . , Xk)be a uniform random variable on the set ofk-tuples of nonnegative integers with even sum n≥2.

(a) For any integerk-vector x withk≥2 we have

P(X≡x mod 2) ≤

n/2+k−1 k−1

n+k−1 k−1

≤exp

− 1

4/k+ 2/n

, (6)

with equality at the first location ifx= 0.

(27)

(b) Consider a partition of{1. . . k}intomnonempty parts so thatk≤m`

for some `≥2. Then with ∧denoting minimum, we have P

 X

i∈p

Xi is even for each part p

≤14 exp

−m∧(n/`) 14

.

Proof. (a) (We thank P. Csikv´ary for this simplification of our previous proof.) To count the number of k tuples that are equal to x mod 2, we subtract 1 from each odd entry and divide each resulting entry by 2. We get a bijection between suchk-tuples and the number ofk-tuples with entry sum (n−o)/2, whereo is the number of odd entries ofx. Thus

P(X≡x mod 2) =

(n−o)/2+k−1 k−1

n+k−1 k−1

n/2+k−1 k−1

n+k−1 k−1

.

This shows the first inequality. For the second, note that the right hand side equals

n/2 + 1 n+ 1

n/2 + 2

n+ 2 · · ·n/2 +k−1 n+k−1 , Each factor is at most 1−n+k−1n/2 , giving a bound of

exp

−(k−1)n/2 n+k−1

≤exp

− 1/2 2/k+ 1/n

.

The last inequality holds fork≥2.

(b) Let ¯X denote the vector formed by the sums of the entries of X over the parts of our partition. LetM ⊂ {1, . . . , k} be a subset of indices, one in each part, and letM0 be its complement. Let S=P

i∈MXi. Then ES= X

i∈M

EXi=mn k ≥ n

`.

We first bound the probability thatS is exceptionally small, namely that it is at most (k∧n)/(4`). S has a discrete beta distribution. By a standard construction, S +m can be realized as the time of the mth black sample when sampling without replacement from n white and k−1 black balls.

From this we get

ps=P(S=s) =

s+m−1 m−1

(n−s)+(k−m)−1 (k−m)−1

n+k−1 k−1

.

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

Preferential attachment rule in a random graph model means, that when a new vertex is born, then the probability that the new vertex will be connected to an old vertex is

Ordered Representations Given a graph G and a permutation π of the vertices, we say that a grounded (segment or string) representation of G is π-ordered if the base points of

The number of edges in the graph is given, determine the maximum number of distances k or paths of length k in the graph, either fixing the number of vertices or not.. Let the

For a given nonincreasing sequence d of nonnegative integers, the following algorithm decides whether d can be realized by simple graph or not.. Moreover, if the answer is yes,

If the graph G has the property that removing any k edges of G, the resulting graph still contains (not necessarily spans) a subgraph isomorphic with H, then we say that G is k

T.L. The bipartite graph K 3,3 is not 2- choosable. Let G be a planar graph and let C be the cycle that is the boundary of the exterior face. We can assume that the planar graph G

Edge Clique Cover : Given a graph G and an integer k, cover the edges of G with at most k cliques.. (the cliques need not be edge disjoint) Equivalently: can G be represented as

Edge Clique Cover : Given a graph G and an integer k, cover the edges of G with at most k cliques. (the cliques need not be edge disjoint) Equivalently: can G be represented as